• Aucun résultat trouvé

Polarizable Force Fields for Biomolecular Modeling

N/A
N/A
Protected

Academic year: 2021

Partager "Polarizable Force Fields for Biomolecular Modeling"

Copied!
70
0
0

Texte intégral

(1)

HAL Id: hal-01114184

https://hal.sorbonne-universite.fr/hal-01114184

Submitted on 3 Nov 2015

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Polarizable Force Fields for Biomolecular Modeling

Yue Shi, Pengyu Ren, Michael Schnieders, Jean-Philip Piquemal

To cite this version:

Yue Shi, Pengyu Ren, Michael Schnieders, Jean-Philip Piquemal. Polarizable Force Fields for Biomolecular Modeling. Reviews in Computational Chemistry, 28 (51), Wiley, 2015, �10.1002/9781118889886.ch2�. �hal-01114184�

(2)

1

Polarizable Force Fields for Biomolecular Modeling

1

2

3

Yue Shi, Pengyu Rena 4

a. Department of Biomedical Engineering 5

The University of Texas at Austin 6 1 University Station, C0800 7 Austin, TX 78712 8 9 10 Michael Schniedersb 11

b. Department of Biomedical Engineering, College of Engineering and 12

Department of Biochemistry, Carver College of Medicine 13

The University of Iowa 14

Iowa City, Iowa, 52242 15

16 17

Jean-Philip Piquemalc 18

c. Laboratoire de Chimie Théorique (UMR 7616), 19

UPMC, Sorbonne Universités, 20

CC 137, 4 place Jussieu, 21

75252 Paris Cedex 05, France 22

(3)

2

1. Introduction

24

Molecular mechanics based modeling has been widely used in the study of chemical and 25

biological systems. The classical potential energy functions and their parameters are 26

referred to as force fields. Empirical force fields for biomolecules emerged in the early 27

1970's,1 followed by the first molecular dynamics simulations of the bovine pancreatic 28

trypsin inhibitors (BPTI).2-4 Over the past 30 years, a great number of empirical 29

molecular mechanics force fields, including AMBER,5 CHARMM,6 GROMOS,7 OPLS,8 30

and many others, have been developed. These force fields share similar functional forms, 31

including valence interactions represented by harmonic oscillators, point dispersion-32

repulsion for van der Waals (vdW) interactions, and an electrostatic contribution based 33

on fixed atomic partial charges. This generation of molecular mechanics force fields has 34

been widely used in the study of molecular structures, dynamics, interactions, design and 35

engineering. We refer interested readers to some recent reviews for detailed discussions.9, 36

10 37

Although the fixed charge force fields enjoyed great success in many areas, there remains 38

much room for improvement. In fixed charge based electrostatic models, the atomic 39

partial charges are meant to be “pre-polarized” for condensed phases in an averaged 40

fashion, typically achieved by the fortuitous overestimation of electrostatic charges by 41

low-level ab initio quantum mechanics. Such models thus lack the ability to describe the 42

variation in electrostatics due to many-body polarization effects, which have been shown 43

to be a significant component of intermolecular forces.10-12 With the rapid growth of 44

computational resources, there has been increasing effort to explicitly incorporate many-45

body induction into molecular mechanics to improve the accuracy of molecular modeling. 46

(4)

3

Classical electrostatics models that take into account polarization appeared as early as the 47

1950s. Barker in his 1953 paper “Statistical Mechanics of Interacting Dipoles” discussed 48

the electrostatic energy of molecules in terms of “permanent and induced dipoles”.13 49

Currently, polarizable models generally fall into three categories: those based on induced 50

point dipoles,9, 14-23 the classical Drude oscillators,24-26 and fluctuating charges.27-30 More 51

sophisticated force fields that are “electronic structure-based”31, 32 or use “machine 52

learning methods”33 also exist, but incur higher computational costs. Discussions of the 53

advantages and disadvantages of each model and their applications will be presented in 54

the following sections. 55

Compared to fixed charge models, the polarizable models are still in a relatively early 56

stage. Only in the past decade or so has there been a systematic effort to develop general 57

polarizable force fields for molecular modeling. A number of reviews have been 58

published to discuss various aspects of polarizable force fields and their development.9, 34-59

40

Here, we focus on the recent development and applications of different polarizable 60

force fields. We begin with a brief introduction to the basic principles and formulae 61

underlying alternative models. Next, the recent progress of several well-developed 62

polarizable force fields is reviewed. Finally, applications of polarizable models to a 63

range of molecular systems, including water and other small molecules, ion solvation, 64

peptides, proteins and lipid systems are presented. 65

1. Modeling Polarization Effects

66

1.1. Induced Dipole Models

(5)

4

To describe electrostatic interactions involving polarization, we consider a system 68

consisting of a collection of charge distribution sites located at lone-pair positions, atomic 69

centers and/or molecular centers, depending on the resolution of the model. The total 70

charge distribution at site i is the sum of permanent and induced charge 71

[1] 72

where M represents the charge distribution. This distribution can be a simple point charge, 73

a point multipole expansion with charge, dipole, quadrupole and/or higher order moments, 74

or a continuous charge distribution. While the principles described below are not limited 75

to any particular representation of charge distribution, we will use point multipoles for 76

convenience. 77

The electrostatic interaction energy between two charge sites i and j is given by 78

[2] 79

where T is the interaction operator and is a function of the distance between i and j. In the 80

case of point charge interactions, T is simply 1/r. The work (positive energy) needed to 81

polarize a charge distribution also has a quadratic dependence on the induced charge 82

distribution: 83

[3] 84

where α is the polarizability of site i that includes all orders of polarizability including 85

dipole polarizability.41 Although α is generally treated as an isotropic quantity, as in the 86

(6)

5

Applequist scheme 41, ab initio anisotropic polarizability tensors can be derived from 87

quantum mechanical calculations.42, 43 88

The total electrostatic energy is 89

[4] 90

The values of the induced moments minimize the total energy, by satisfying 91 [5] 92 As a result 93 [6] 94

Equation [6] can be solved iteratively to obtain the induced dipoles. The self-consistent 95

calculation is computationally expensive; however it can be accelerated with predictors 96

and non-stationary iterative methods.44 97

Substituting from Eq [5] into Eq [6], the final electrostatic energy becomes 98

[7]

99

where the first term is the permanent electrostatic energy and the second term is the 100

polarization energy. 101

1.2. Classic Drude Oscillators

102

In the Drude oscillator model, the polarization effect is described by a point charge (the 103

Drude oscillator) attached to each non-hydrogen atom via a harmonic spring. The point 104

(7)

6

charge can move relative to the attachment site in response to the electrostatic 105

environment. The electrostatic energy is the sum of the pairwise interactions between 106

atomic charges and the partial charge of the Drude particles 107

, 108

[8] 109

where ND and N are the number of Drude particles and non-hydrogen atoms, qD and qC

110

are the charges on the Drude particle and its parent atom, respectively, rD and rC are their

111

respective positions, and kD is the force constant of the harmonic spring between the

112

Drude oscillator and its parent atom. The last term in Equation [8] accounts for the cost 113

of polarizing the Drude particles. 114

The atomic polarizability (α) is a function of both the partial charge on the Drude particle 115

and the force constant of the spring 116

[9] 117

Both the induced-dipole and Drude oscillator approaches benefit from short-range Thole 118

damping to avoid a polarization catastrophe and to produce an anisotropic molecular 119

polarization response.45 120

1.3. Fluctuating Charges

(8)

7

The formalism of the fluctuating charge model is based on the charge equilibration 122

(CHEQ) method,46 in which the chemical potential is equilibrated via the redistribution of 123

charge density. The charge-dependent energy for a system of M molecules containing Ni

124

atoms per molecule is expressed as 125

126

(10) 127

where Qi is the partial charge on atomic site i. The χ describes the atomic

128

electronegativity controlling the directionality of electron flow, and J is the atomic 129

hardness that represents the resistance to electron flow to or from the atom. These 130

parameters are optimized to reproduce molecular dipoles and the molecular polarization 131

response. The charge degrees of freedom are typically propagated via an extended 132

Lagrangian formulation:47 133

[11] 134

where the first two terms represent the nuclear and charge kinetic energies, the third term 135

is the potential energy, and the fourth term is the molecular charge neutrality constraint 136

enforced on each molecule i via a Lagrange multiplier λi. The extended Lagrangian

137

approach can also be applied to the induced dipole and Drude oscillator models described 138

earlier. While the extended Lagrangian seems to be more efficient than the iterative 139

method, fictitious masses and smaller time-steps are required to minimize the coupling 140

(9)

8

between the polarization and atomic degrees of freedom, which can never be completely 141

eliminated.44 142

A few general force fields have been developed based on these formulas to explicitly 143

treat the polarization effect. We now discuss development highlights for some of the 144

representative force fields. 145

2. Recent Developments

146

2.1. AMOEBA

147

The AMOEBA (Atomic Multipole Optimized Energetics for Biomolecular Applications) 148

force field, developed by Ponder, Ren and co-workers,15, 18, 37 utilizes atomic multipoles 149

to represent permanent electrostatics and induced atomic dipoles for many-body 150

polarization. The valence interactions include bond, angle, torsion and out-of-plane 151

contributions using typical molecular mechanics functional forms. The van der Waals 152

interaction is described by a buffered-14-7 function. The atomic multipole moments 153

consist of charge, dipole and quadrupole moments, which are derived from ab initio 154

quantum mechanical calculations using procedures such as Stone’s Distributed Multipole 155

Analysis (DMA).48-50 The higher order moments make possible anisotropic 156

representations of the electrostatic potential outside atoms and molecules. The 157

polarization effect is explicitly taken into account via atomic dipole induction. The 158

combination of permanent atomic multipoles and induced dipoles enables AMOEBA to 159

capture electrostatic interactions in both gas and condensed phase accurately. The vdW 160

parameters of AMOEBA are optimized simultaneously against both ab initio gas-phase 161

data and condensed-phase experimental properties. 162

(10)

9

In the past decade, AMOEBA has been applied to the study of water,15 monovalent and 163

divalent ions,51-53 small molecules,54, 55 peptides18, 56 and proteins.57-59 AMOEBA 164

demonstrated that a polarizable force field is able to perform well in both gas and 165

solution phases with a single set of parameters. In addition, AMOEBA is the first 166

general-purpose polarizable force field utilized in molecular dynamics simulations of 167

protein-ligand binding and calculation of absolute and relative binding free energies.58-62 168

The computed binding free energies between trypsin and benzamidine derivatives 169

suggests significant non-additive electrostatic interactions as the ligand desolvates from 170

water and enters the protein pocket (see Section 4.4 for further discussion). AMOEBA 171

has recently been extended to biomolecular X-ray crystallography refinement63, 64, and 172

consistently successful prediction of the structure, thermodynamic stability and solubility 173

of organic crystals65 are encouraging. 174

AMOEBA has been implemented in several widely used software packages including 175

TINKER,66 OpenMM,67 Amber,68 and Force Field X.69 The AMOEBA polarizable force 176

field was first implemented within the FORTRAN-based TINKER software package70 177

using Particle Mesh Ewald (PME) for long-range electrostatics. Implementation of the 178

polarizable-multipole Poisson-Boltzmann,71 which depends on the Adaptive Poisson-179

Boltzmann Solver (APBS),72 and generalized Kirkwood73 continuum electrostatics 180

models also exist in TINKER, which is now being parallelized using OpenMP. The 181

algorithms in TINKER are also available from within CHARMM using the MSCALE 182

interface.74, 75 Alternative FORTRAN implementations of AMOEBA using PME are 183

available in the Sander and PMEMD molecular dynamics engines of AMBER,68 with the 184

latter parallelized using MPI. The PME treatment of AMOEBA electrostatics has recently 185

(11)

10

been extended within the Java Runtime Environment (JRE) program Force Field X by 186

incorporating explicit support for crystal space group symmetry,63 parallelizing for 187

heterogeneous computer hardware environments63 and supporting advanced free energy 188

methods such as the Orthogonal Space Random Walk (OSRW) strategy.65, 76 These 189

advancements are critical for applications such as AMOEBA-assisted biomolecular X-ray 190

refinement,63, 77 efficient computation of protein-ligand binding affinity,57, 61 and 191

prediction of the structure, stability and solubility of organic crystals.65 Finally, the 192

OpenMM software is working toward a general implementation of AMOEBA using the 193

CUDA GPU programming language.78 194

2.2. SIBFA

195

The SIBFA (Sum of Interactions Between Fragments Ab initio computed) force field for 196

small molecules and flexible proteins, developed by Gresh, Piquemal et. al,79-83 is one of 197

the most sophisticated polarizable force fields because it incorporates polarization, 198

electrostatic penetration 84 and charge-transfer effects.85 199

The polarization is treated with an induced dipole model, in which the distributed 200

anisotropic polarizability tensors43 are placed on the bond centers and on the heteroatom 201

lone pairs. Quadrupolar polarizabilities are used to treat metal centers. The force field is 202

designed to enable the simultaneous and reliable computation of both intermolecular and 203

conformational energies governing the binding specificities of biologically and 204

pharmacologically relevant molecules. Similar to AMOEBA, permanent multipoles are 205

used for permanent electrostatics in SIBFA. Flexible molecules are modeled by 206

combining the constitutive rigid fragments. SIBFA is formulated on the basis of quantum 207

(12)

11

chemistry and calibrated on energy decomposition analysis, as oppose to AMOEBA 208

which relies more on condensed-phase experimental data. It aims to produce accurate 209

interaction energy comparable with ab initio results. The development of SIBFA 210

emphasizes separability, anisotropy, nonadditivity and transferability. The analytical 211

gradients for charge-transfer energy and solvation contribution are not yet available in 212

SIBFA although molecular dynamics simulations with a simplified potential have been 213

attempted and will be reported in the near future. 214

SIBFA has been validated on a wide range of molecular systems from water clusters86 to 215

large complexes like metalloenzymes encompassing Zn(II).87-92 It has been used to 216

investigate molecular recognition problems including the binding of nucleic acids to 217

metal ions,93-95 the prediction of oligopeptide conformations,86, 96 and for ligand-protein 218

binding.97 Most of the SIBFA calculations reproduced closely the quantum chemistry 219

results, including both the interaction energy and the decomposed energy terms. At the 220

same time, electrostatic parameters are demonstrated to be transferable between similar 221

molecules. 222

,A Gaussian based electrostatic model (GEM) has been explored as an alternative to 223

distributed point multipole electrostatic representation.98 GEM computes the molecular 224

interaction energies using an approach similar to SIBFA but replacing distributed 225

multipoles by electron densities.99 GEM better captures the short-range effects on 226

intermolecular interaction energies, and it naturally includes the penetration effect. 227

Calculations on a few simple systems like water clusters99 have demonstrated GEM’s 228

capability to reproduce quantum chemistry results. Furthermore, implementating PME 229

for GEM in a PBC showed reasonable computational efficiency thanks to the use of 230

(13)

12

Hermite Gaussian functions.100 Therefore, replacing SIBFA’s distributed multipoles with 231

the GEM continuous electrostatic model will be a future direction of methodology 232

development.98 233

2.3. NEMO

234

NEMO (Non-Empirical Molecular Orbital) is a polarizable potential developed by 235

Karlström and co-workers.101-103 The NEMO potential energy function is composed of 236

electrostatics, induction, dispersion and repulsion terms. The induction component is 237

modeled using induced point–dipole moments with recent addition of induced point– 238

quadrupole moments.22 The electrostatics, previously represented by atomic charges and 239

dipoles, has also been extended to include atomic quadrupole moments leading to notable 240

improvement on formaldehyde. The atomic multipole moments are now obtained from ab 241

initio calculation using a LoProp procedure.104 The LoProp is claimed to provide atomic 242

multipoles and atomic polarizabilities that are less sensitive to basis sets than are other 243

methods such as Distributed Multipole Analysis (DMA). Also, NEMO is the only force 244

field that explores the possibility of including interactions between permanent multipoles 245

and higher-order induced multipoles involving higher-order hyperpolarizabilities.22 246

NEMO has demonstrated its ability to describe accurately both inter and intramolecular 247

interactions in small systems, including: glycine dipeptide conformation profiles,105 ion-248

water droplets,106 and urea transition from nonplanar to planar conformation in water.107 249

Its applicability to biomacromolecules is not yet known. 250

2.4. CHARMM-Drude

(14)

13

In addition to the induced dipole model, the classical Drude oscillator model is another 252

popular approach for modeling polarization effects.39, 108 Roux, MacKerell and their 253

colleagues have been developing a polarizable CHARMM force field based on this 254

approach. 25, 26, 109, 117 Unlike the induced dipole model, which treats the polarization 255

response using point dipoles, the Drude model represents the polarizable centers by a pair 256

of point charges. A point partial charge is tethered via a harmonic spring for each non-257

hydrogen atom. This point charge (the Drude oscillator) can react to the electrostatic 258

environment and cause the displacement of the local electron density. The atomic 259

polarizability depends on both the Drude particle charge and the harmonic force constant. 260

In MD simulations, the extended Lagrangian is used to evaluate the polarization response, 261

by allowing the Drude particles to move dynamically and experience nonzero forces. 262

Small fictitious masses are assigned to each Drude particle and independent low 263

temperature thermostats are applied to the Drude particle degrees of freedom.118 In case 264

of energy minimization, self-consistent iteration will be required to solve for the 265

polarization. 266

Determining electrostatic parameters for the Drude oscillator is not as straightforward as 267

for induced dipole models. Masses assigned to the Drude particles are chosen empirically. 268

The values for atomic charges and polarizabilities requires a series of calculations of 269

perturbed ESP maps. This force field has been parameterized for water25, 26, and for a 270

series of organic molecules including: alkanes,110 alcohols,111 aromatics,112 ethers,113, 114 271

amides,109 sulfurs,115 and ions.119, 120 An attempt has also been made to combine the 272

Drude-based polarizable force field with quantum mechanics in QM/MM methods.121 It 273

was noted that pair-specific vdW parameters are needed to obtain accurate hydration free 274

(15)

14

energies of small molecules using the polarizable force field. This is likely due to the 275

problematic combining rules used to compute the vdW interactions between unlike atoms. 276

The Drude model has been implemented in CHARMM74, 122 and in the NAMD 277

package,123 in which the computational cost is about 1.2 to 1.8 times greater than that of 278

fixed-charge CHARMM.124 279

2.5. CHARMM-FQ

280

The fluctuating charge model (FQ), also known as charge equilibration or 281

electronegativity equalization model, is an empirical approach for calculating charge 282

distributions in molecules. In this formalism, the partial charge on each atom is allowed 283

to change to adapt to different electrostatic environments. The variable partial charges are 284

computed by minimizing the electrostatic energy for a given molecular geometry. 285

Compared with the induced dipole and Drude models, the fluctuating charge models are 286

minimally parameterized and easier to implement because the polarizability is induced 287

without introducing new interactions beyond the point charges. Either extended 288

Lagrangian or self-consistent iteration can be used to compute the fluctuating charges in 289

MD simulations, with similar advantages and disadvantages as discussed above. 290

The CHARMM-FQ force field,125, 126 developed by Patel, Brooks, and their coworkers, 291

has been parameterized for small molecules,28 proteins,28, 127 lipids, lipid bilayers,113, 128 292

and carbohydrates.125 The force field has been applied to investigate liquid–vapor 293

interfaces in addition to biophysical studies.129 There are some known limitations for 294

fluctuating charge models, however, such models allow artificial charge transfer between 295

widely separated atoms but that can be controlled with additional constraints. Also the 296

(16)

15

intramolecular charge-flow is limited by the chemical connectivity. It is thus difficult to 297

capture the out-of-plane polarization in molecules such as aromatic benzenes with 298

additional charge sites. The CHARMM-FQ force field has been implemented in the 299

CHARMM software package.74 300

2.6. X-Pol

301

Gao and coworkers proposed the X-Pol framework by combining the fragment-based 302

electronic structure theory with a molecular mechanical force field.31, 32, 130 Unlike the 303

traditional force fields, X-Pol does not require bond stretching, angle, and torsion terms 304

because they are represented explicitly by quantum mechanics. The polarization and 305

charge transfer between fragments are also evaluated quantum mechanically.130 306

Furthermore, X-Pol can be used to model chemical reactions. 307

In X-Pol, large molecular systems are divided into small fragments. Electrostatic 308

interactions within the fragments are treated using the electronic structure theory. The 309

electrostatic interactions between fragments are described by the combined quantum 310

mechanical and molecular mechanical (QM/MM) approach. Also, a vdW term is added to 311

the interfragment interaction as a consequence of omitting electron correlation and 312

exchange repulsion. A double self-consistent-field (DSCF) procedure is used to converge 313

the total electronic energy of the system as well as the energy within the fragments (this 314

includes the mutual polarization effect). 315

The X-Pol potential has been applied to MD simulations of liquid water,131 liquid 316

hydrogen fluoride,132 and covalently bonded fragments.133, 134 This model was recently 317

used in a molecular dynamics simulation of a solvated protein.135 As expected the 318

(17)

16

computational efficiency of the X-Pol is in between that of a simple classical force field 319

and a full ab initio method. The solvated trypsin required 62.6 h to run a 5 ps simulation 320

on a single 1.5 GHz IBM Power4 processor. A parallel version of X-Pol is being 321

developed. 322

2.7. PFF

323

Kaminski et al. developed a polarizable protein force field (PFF) based on ab initio 324

quantum theory.136, 137 The electrostatic interaction is modeled with induced dipoles and 325

permanent point charges. With the exception of a dispersion parameter, all other 326

parameters, including the electrostatic charges and polarizabilities, are obtained by fitting 327

to quantum chemical binding energy calculations for homodimers. The dispersion 328

parameters are later refined by fitting to the experimental densities of organic liquids.16 329

Gas-phase many-body effects, as well as conformational energies, are well reproduced,137 330

and MD simulations for real proteins are reasonably accurate at modest computational 331

costs.16, 138 332

To reduce the computational cost, a POSSIM (Polarizable Simulations with Second-order 333

Interaction Model) force field was later proposed, in which the calculation of induced 334

dipoles stops after one iteration.139, 140 The computational efficiency can be improved by 335

almost an order of magnitude by using this formalism. Because the analytical gradients 336

(forces) are unavailable, a Monte-Carlo technique is used in condensed-phase simulations. 337

POSSIM has been validated on selected small model systems, showing good agreement 338

with ab initio quantum mechanical and experimental data. Parameters for alanine and 339

protein backbone have been reported.141 340

(18)

17

Polarizable force fields for non-biological systems also exist. A many-body polarizable 341

force field by Smith and coworkers was developed and applied to the simulations of ion 342

conduction in polyethylene oxide (PEO).142-144 Cummings and coworkers developed an 343

interesting Gaussian charge polarizable force field for ions and in polyethylene oxide 344

(PEO).145-147 A polarizable force field for ionic liquids was also reported to provide 345

accurate thermodynamics and transport properties.148 346

3. Applications

347

3.1. Water Simulations

348

Due to its important role in life, water is a natural choice for polarizable force field 349

development. After the polarizable (and dissociable) water model of Stillinger and 350

David,149 more than a dozen polarizable water models have been reported.150 351

Similar to how the polarization models discussed previously, the polarizable water 352

models likewise fall into three major categories. Most belong to the first category, 353

including the Stillinger and David’s water model, SPCP,151 PTIP4P, 152 CKL,153 NCC,154 354

PROL,155 Dang-Chang156 and others. These models all adopted the induced dipole 355

framework to treat polarization, typically using a single polarizable site on water. TTM 356

models157-160 and the AMOEBA water model15 utilize an interactive, distributed atomic 357

polarizability with Thole’s damping scheme45 to treat electrostatics and polarization. The 358

Drude Oscillator-based water models include SWM4-DP,26 and SWM4-NDP,25 as well as 359

the Charge-On-Spring (COS) model,161 and its improved variation.162 The third group 360

includes the SPC-FQ and TIP4P-FQ163 water models that utilize the fluctuating charge 361

scheme to model polarization. The partial charges flow from one atom to another, and the 362

(19)

18

total charge of a water molecule need not be zero. Stern et al. proposed a unique water 363

model (POL5) by combining the fluctuating charge with the point induced dipole 364

scheme.164 Several more sophisticated polarizable water models based on quantum 365

mechanics were developed based on quantum mechanics, including QMPFF,165 DPP2,166 366

and Polarflex.167 For example, the charge penetration, induction, and charge transfer 367

effects have been incorporated into the DPP2 (Distributed Point Polarizable Model) 368

model which reproduces well the high-level ab initio energetics and structures for large 369

water clusters. 370

An advantage of a polarizable water model over most non-polarizable models is the 371

ability to describe the structure and energetics of water in both gas and condensed phases. 372

Water dimer interaction energies, the geometry of water clusters and the heat of 373

vaporization of neat water can be reproduced well by most polarizable models. Some 374

highly parameterized nonpolarizable force fields such as TIP5P, TIP4P-EW and 375

TIP4P/2005 actually perform as well or better than some polarizable force fields over a 376

range of liquid properties, including the density-temperature profile, radial distribution 377

function, and diffusion coefficient. However, for water molecules experiencing 378

significant changes in environment, e.g., from bulk water to the vicinity of ions or 379

nonpolar molecules, only the polarizable models can capture the change of water dipole, 380

structure and energetics.168 381

Polarization water models are being extended and applied to other phases as well as to 382

the interface between different phases. Rick et al recently incorporated charge transfer 383

into their polarizable water model that was then used to study ice/water coexistence 384

properties and properties of the ice Ih phase.169 The POL3 water model14, 170 was used to 385

(20)

19

study the ice-vapor interface, and to calculate the melting point of ice Ih. Bauer and Patel 386

used the TIP4P-QP model to study the liquid-vapor coexistence.171 387

3.2. Ion Solvation

388

Ions are an important component in many chemical and biological systems. Nearly half 389

of all proteins contain metal ions, and they play essential roles in many fundamental 390

biological functions. Some metal ions are critical for both protein structure and function. 391

In enzymes, ions can bind and orient the substrates through electrostatic interactions at 392

the active sites, thus controlling catalytic reaction. Divalent ions are vital in nucleic acid 393

structures. Modeling ion-water and ion-biomolecule interactions accurately is very 394

important. 395

Due to the high electron density and small sizes of ions, the non-polarizable models fail 396

to capture the structural details adequately and do not or to reproduce the atomic dipole 397

of water around the ions.172-176 Several studies of ion solvation have been reported using 398

different polarizable models51-53, 116, 120, 177-187 with analyses focused on solvation 399

structures, charge distribution, and binding energies. Noted that no straightforward 400

experimental measurement of hydration free energy data exist because the macroscopic 401

system must be neutral. Different assumptions are used to decompose the experimental 402

hydration free energy into single ion contributions. The hydration free energy of some 403

monovalent ions such as Na+ and K+ from different sources can vary by as much as 10 404

kcal/mol. It is more reliable to compare the hydration free energy of the whole salt and 405

the relative energy between cations or anions. 406

(21)

20

The AMOEBA polarizable force field has been used to model a number of anions and 407

cations, including Na+, K+, Mg++, Ca++, Zn++, Cl-, Br-, and I-.51-53, 188 Parameters for these 408

ions, including the vdW parameters and polarization damping coefficients (for divalent 409

ions only), were obtained by fitting to the ab initio QM interaction energy profiles of ion-410

water pairs. Molecular dynamics simulations were then performed to evaluate the ion-411

cluster solvation enthalpies and solvation free energies.51-53, 188 The excellent agreement 412

between calculated and experimental hydration free energy, often within 1%, demonstrate 413

that polarizable force fields are transferable between phases. Ab initio energy 414

decomposition using, e.g., the Constrained Space Orbital Variations (CSOV) method,99, 415

189

have also been applied to examine the polarization component of the ion-water 416

interaction energy and to guide the force field parameterization.53, 190 More recently, the 417

AMOEBA force field was used to model the hydration of high valent Th(IV)94 and 418

studies on open-shell actinides are underway. 419

The SIBFA model was used to examine Pb(II),191 lanthanides (La(III) and Lu(III)) and 420

actinides (Th(IV)) in water.94 SIBFA-predicted interaction energies generally matched 421

well with the ab initio results, including the energy decompositions. Lamoureux and 422

Roux developed the CHARMM polarizable force field for alkali and halide ions based on 423

the Drude Oscillator.177 Hydration free energies, calculated via thermodynamic 424

integration,192 showed an encouraging agreement with experiment. 425

3.3. Small Molecules

426

Small molecules are building blocks of biomolecules and serve as substrates and 427

inhibitors. Abundant experimental measurements on various physical and chemical 428

(22)

21

properties exist for common organic molecules which in turn are used in the 429

parameterization of the force fields. Polarizable and non-polarizable force fields can 430

usually produce reasonable estimations of physical properties of neat liquids.193-196 431

Extensive studies using polarizable force fields, covering major functional group, 432

including alkanes, alcohols, aldehydes, ketones, ethers, acids, aromatic compounds, 433

amines, amides, and some halogen compounds have been reported.28, 36, 55, 110, 112, 126, 197-434

199

Calculations of structure, dipole moment, heterodimer binding energy, liquid diffusion 435

constant, density, heat of vaporization, and hydration free energy are usually performed 436

to assess the quality of force field parameters. 437

The electrostatic multipole parameters in AMOEBA were derived using the DMA 438

procedure. They can be further optimized to the electrostatic potentials of chosen ab 439

initio theory and basis sets. The AMOEBA valence parameters were derived from ab

440

initio data such as molecular geometries and vibrational frequencies of the gas-phase

441

monomer. The vdW parameters are estimated from gas-phase cluster calculations, and 442

subsequently refined in liquid simulations using experimental data (e.g., densities and 443

heats of vaporization). The torsional parameters the last obtained during the 444

parameterization scheme are derived by fitting to ab initio QM conformational energy 445

profiles. An automated protocol (PolType) that can generate AMOEBA parameters for 446

small molecules is under development.200 Because force field parameterization is a 447

tedious process, such an automated tool is convenient and reduces the likelihood of 448

human error. 449

The CHARMM-Drude force field developers devoted much of their efforts on organic 450

compounds. Their parameterization scheme starts from an initial guess of charge (based 451

(23)

22

on the CHARMM22 force field), and invokes changes at some lone pair sites. Those 452

parameters are then fit to a series of “perturbed” ESP maps. The vdW parameters are then 453

optimized to match neat liquid properties as is done many other force fields.115 Overall, a 454

systematic improvement over the CHARMM22 additive force field has been observed for 455

both gas-phase and condensed-phase properties. These studies on small molecules lay the 456

groundwork for developing a Drude-based polarizable force field for proteins and nucleic 457

acids. 458

3.4. Proteins

459

One of the goals for polarizable force fields is to model accurately protein structures, 460

dynamics, and interactions. Proteins are a ubiquitous class of biopolymers whose 461

functionalities depend on the details of their 3D structures, which, in turn, are largely 462

determined by their amino acid sequences. Fixed-charge force fields for proteins, like 463

AMBER, CHARMM, and OPLS-AA, have been developed and for years subjected to 464

various tests and validations. The development of polarizable protein force fields is still 465

in its infancy. Although the importance of including polarization effects was recognized 466

long ago, polarizable protein force fields emerged only in the past decade.9, 21, 28, 29, 37, 138, 467

201-205 468

The use of polarizable electrostatics in protein simulations dates back to 1976,1 when 469

Warshel and Levitt simulated lysozyme via single point calculations. Kaminski et al. 470

reported in 2002 an ab initio polarizable protein force field (PFF) based on inducible 471

dipoles and point charges.16,137 Simulations on bovine pancreatic trypsin inhibitor using 472

PDFF showed a satisfactory root mean square displacement (RMSD) compared to the 473

(24)

23

experimental crystal structure and polarization was found to affect the solvation 474

dynamics.138 The fluctuating-charge based ABEEM/MM force field was used to examine 475

protein systems like trypsin inhibitors206 and the heme prosthetic group.207 The SIBFA 476

force field has been used to study the interaction between focal adhesion kinase (FAK) 477

and five pyrrolopyrimidine inhibitors.208 The energy balances accounting for the 478

solvation/desolvation effects calculated by SIBFA agree with experimental ordering. 479

Water networks in the binding pocket were shown to be critical in terms of binding 480

affinity. Moreover, the polarization contribution was considered as an indispensable 481

component during the molecular recognition. In comparison, the continuum reaction field 482

procedure fails to reproduce these properties. In addition kinases, the SIBFA protein 483

force field has been used to study a variety of metalloproteins encompassing cations such 484

as Cu+, Zn++, Ca++ or Mg++, as well as enabling inhibition studies.91, 209-211 Future 485

molecular dynamics simulations should extend the applicability of SIBFA to protein-486

ligand binding. 487

Ren and coworkers have been systematically developing the AMOEBA protein force 488

field, and using it to study to several protein systems to understand protein-ligand 489

binding.57-59, 61 More recently an X-Pol force field for proteins has been developed and 490

demonstrated in a simulation of solvated trypsin.32 491

The first attempt to compute the protein-ligand binding free energy using a polarizable 492

force field was made on the trypsin-benzamidine systems using AMOEBA.57, 61, 62 The 493

absolute binding free energy of benzamidine to trypsin, and the relative binding free 494

energies for a series of benzamidine analogs, were computed using a rigorous alchemical 495

transformation. AMOEBA was successful in evaluating the binding free energies 496

(25)

24

accurately with an average error well within 1.0 kcal/mol. A similar study on trypsin, 497

thrombin and urokinase was reported using another ab initio QM-based polarizable force 498

field.212 A thermodynamic integration scheme was used to compute the relative binding 499

free energies, which were in excellent agreement with experimental data (root mean 500

squre error (RMSE)=1.0 kcal/mol). 501

AMOEBA was later used to examine an “entropic paradox” associated with ligand 502

preorganization discovered in a previous study of conformationally constrained 503

phosphorylated-peptide analogs that bind to the SH2 domain of the growth receptor 504

binding protein 2 (Grb2).59 The paradox refers to the unusual trend in which the binding 505

of unconstrained peptides (assumed to lose more entropy upon binding) is actually more 506

favorable entropically than are the constrained counterparts. AMOEBA correctly 507

reproduced the experimental trend and at the same time repeated a mechanism in which 508

the unconstrained peptide ligands were “locked” by intramolecular nonbonded 509

interactions. The simulations uncovered a crucial caveat that had not been previously 510

acknowledged regarding the general design principle of ligand preorganization, which is 511

presumed by many to have a favorable effect on binding entropy. 512

More recently, Zhang et al. demonstrated the ability of AMOEBA in dealing with 513

systems with a metal ion.58 Those authors studied the Zinc-containing matrix 514

metalloproteinases (MMPs) in a complex with an inhibitor where the coordination of 515

Zn++ waswith organic compounds and protein side chains. Polarization was found to play 516

a key role in Zn++ coordination geometry in MMP. In addition, the relative binding free 517

energies of selected inhibitors binding with MMP13 were found to be in excellent 518

agreement with experimental results. As with the previous trypsin study, it was found that 519

(26)

25

binding affinities are likely to be overestimated when the polarization between ligands 520

and environments is ignored. 521

Having a more rigorous physical model for treating polarization, the ability to model 522

protein-ligand interactions has been improved significantly. Systems involving highly 523

charged species, like metal ions, can now be treated with confidence. This in turn, 524

provides tremendous opportunities for investifating important proteins for drug discovery 525

and for protein engineering. 526

3.5. Lipids

527

With the rapid development of computational resources, simulations of large systems like 528

lipid bilayers with membrane proteins is feasible.126, 213 Patel and coworkers have been 529

developing a polarizable force field for biomembranes to study the structure and 530

dynamics of ion channel systems.40, 113, 128, 214 Simulations of solvated DMPC 531

(dimyristoyl phosphatidylcholine) and dipalmitoylphosphatidylcholine (DPPC) bilayers 532

were reported.113, 214 The distribution of the membrane components along the lipid bilayer 533

is similar to that from a fixed charge model. The water dipole moment was found to 534

increase from about 1.9 Debye in the middle of the membrane plane to the average bulk 535

value of 2.5~2.6 Debye. The lipid surface computed with the polarizable force field was 536

not improved from those of non-polarizable ones however. In addition, ion permeation in 537

a gramicidin A channel embedded in a DMPC bilayer was investigated.113 Davis and 538

Patel concluded that including the electronic polarization lowered the ion permeation free 539

energy barrier significantly, from 12 kcal/mol to 6 kcal/mol. 540

3.6. Continuum Solvents for Polarizable Biomolecular Solutes

(27)

26

A continuum solvent replaces explicit atomic details with a bulk, mean-field response. It 542

is possible to demonstrate from statistical mechanics that an implicit solvent potential of 543

mean force (PMF) exists, wihch preserves exactly the solute thermodynamic properties 544

obtained from explicit solvent.215 It is possible to formulate a perfect implicit solvent in 545

principle, but in practice approximations are necessary to achieve efficiency. This 546

remains an active area of research.216 An implicit solvent PMF can be formulated via a 547

thermodynamic cycle that discharges the solute in vapor, grows the uncharged (apolar) 548

solute into a solvent and finally recharges the solute within a continuum 549

dielectric 550

[12] 551

The continuum electrostatic energy, including mobile electrolytes, can be described by 552

either the nonlinear Poisson-Boltzmann Equation (NPBE) or the simplified linearized 553

Poisson-Boltzmann Equation (LPBE) 554

[13] 555

where the coefficients are a function of position r, is the potential, is the pemittivity, 556

is the modified Debye-Hückel screening factor, and is the solute charge density.217, 557

218

Implementations of a Poisson-Boltzmann continuum for many-body quantum 558

mechanical potentials have been applied to small molecules for decades. Examples 559

include the Polarizable Continuum Model (PCM) 219, 220, COSMO 221 and the Solvent 560

Model series (SMx).222 In contrast, applications of biomolecular continuum electrostatics 561

have been limited mainly to fixed partial charge solute descriptions for reasons of 562

computing efficiency force field availability. However, as a result of increasing 563

(28)

27

computational power and the completion of the polarizable force fields for biomolecules 564

described above, the coupling of classical many-body potentials to continuum 565

electrostatics is now possible. 566

An important initial demonstration of polarizable biomolecules within a Poisson-567

Boltzmann continuum used the Polarizable Force Field (PFF) of Maple et al. to model 568

protein-ligand interactions.223 A second demonstration used the Electronic Polarization 569

from Internal Continuum (EPIC), which accounts for intramolecular polarization using a 570

continuum dielectric.224, 225 Finally, the polarizable multipole Poisson-Boltzmann (PMPB) 571

model based on the AMOEBA force field demonstrated that the self-consistent reaction 572

field (SCRF) of proteins within a continuum solvent is consonant with the ensemble 573

average response of explicit solvent.71 Contrarily, end-state calculations of protein-ligand 574

binding affinity using the PMPB model were shown to not recapitulate explicit solvent 575

alchemical free energies to chemical accuracy.61 This motivates development of analytic 576

continuum electrostatics (discussed next), which are fast enough to allow binding 577

affinities to be computed using alchemical sampling, rather than merely relying on end-578

states. A key advantage of EPIC is that the biomolecular self-consistent field (SCF) is 579

determined by a single numerical finite-difference (FD) solution of the PBE, unlike the 580

aforementioned atom-centered PFF and PMPB models that require a new solution for 581

each SCF iteration. However, a tradeoff of EPIC’s efficiency gain is a reduction in model 582

flexibility because electrostatic masking rules cannot be incorporated into the FD solver 583

(i.e., the permanent field due to 1-2 or 1-3 interactions cannot be neglected). Although 584

masking of short-range bonded interactions is the standard approach used by essentially 585

all biomolecular force fields, this is not possible for an EPIC style energy model. 586

(29)

28

The first example of an analytic continuum electrostatic model for polarizable 587

biomolecules is the generalized Kirkwood (GK) model for the AMOEBA force field.73 588

The AMOEBA/GK approach has been combined with alchemical sampling to predict 589

trypsin-ligand binding affinity with a correlation coefficient of 0.93. This is a significant 590

improvement over the PMPB end-state approach.226 A second example, based on the 591

ABEEMσπ fluctuating charge force field combined with a generalized Born (GB)

592

continuum electrostatic model, showed promising results for the computation of solvation 593

free energies for small organic molecules and peptide fragments.227 594

3.7. Macromolecular X-ray Crystallography Refinement

595

X-ray crystallography is the dominant experimental method for determining the 3-596

dimensional coordinates of macromolecules. Collected diffraction data is the Fourier 597

transform of the ensemble average electron density of the macromolecular crystal. While 598

reciprocal space amplitudes of Bragg diffraction peaks are measured, their phases are not. 599

Instead, phase information is derived from the Fourier transform of a model structure that 600

is sufficiently close to the actual experimental ensemble. This is known as molecular 601

replacement (MR). After an initial model has been built into the electron density, further 602

refinement is based optimizating a target function of the form 603

[14] 604

where evaluates the agreement between measured and calculated diffraction 605

amplitudes, restrains the model using prior knowledge of intra- and 606

intermolecular chemical forces and weights the relative strength of the two terms.77, 607

(30)

29 228

We now focus on the evolution of the prior chemical knowledge used during the X-608

ray refinement process, and we culminate in ongoing work using polarizable force fields 609

in combination with PME electrostatics algorithms to obtain the most accurate, 610

informative biomolecular models possible. 611

The first application of molecular mechanics to macromolecular X-ray crystallography 612

refinement (based on fixed partial charge electrostatics evaluated using a spherical cutoff) 613

was on influenza-virus hemagglutinin by Weis et al. in 1990.229 This work demonstrated 614

that electrostatics maintained chemically reasonable hydrogen-bonding, although charged 615

surface residues were sometimes observed to form incorrect salt bridges.229 The latter 616

observation highlights the importance of accounting for dielectric screening arising from 617

the heterogeneous distribution of solvent within a macromolecular crystal, by using one 618

of the above described continuum electrostatics models. For example, the generalized 619

Born (GB) model for fixed charge electrostatics has been described, albeit with a 620

spherical cutoff approximation.230 Comparing refinements with and without GB 621

screening showed that roughly 10% of the amino acid side-chain conformations were 622

altered, with 75% of these side-chain differences due to residues at the macromolecular 623

surface.230 Although these first applications of fixed charge force field electrostatics were 624

encouraging, the use of spherical cutoffs to approximate crystal lattice sums is now 625

known to be only conditionally convergent and therefore prone to a variety of artifacts.231 626

In 1921, Ewald introduced an absolutely convergent solution to the problem of evaluating 627

electrostatic lattice summations in crystals. He did this by separating the problem into a 628

short-ranged real space sum and a periodic, smoothly varying, long-range sum that can be 629

evaluated efficiently in reciprocal space.232 This approach, now known as Ewald 630

(31)

30

summation, has been described for both fixed partial charges and atomic multipoles.233, 631

234

More recently, the efficiency of Ewald summation was improved via the particle-mesh 632

Ewald (PME) algorithm, wherein the reciprocal space summation leverages the fast 633

Fourier transforms (FFT)235 via b-Spline interpolation236 for both fixed partial charge and 634

atomic multipole descriptions.237 635

The speed of the PME algorithm has been further improved for crystals by incorporating 636

explicit support for space group symmetry and by parallelization for heterogeneous 637

computer architectures.63 Combining the polarizable AMOEBA force field with 638

electrostatics evaluated using PME has been shown to improve macromolecular models 639

from X-ray crystallography refinement in a variety of contexts.64, 77, 238-240 At high 640

resolution (~1 Å or lower), the information contained within a polarizable atomic 641

multipole force field can be used to formulate the electron density of the scattering model 642

( ), in addition to contributing chemical restraints ( ).64, 238 The importance 643

of the prior chemical information contained in a polarizable force field is most significant 644

when positioning parts of the model that are not discernable from the experimental 645

electron density, as in the orientation of water hydrogen atoms239 or secondary structure 646

elements for mid-to-low resolution data sets (~3-4 Å).63 647

Let use consider an example, the AMOEBA-assisted biomolecular X-ray refinement with 648

electrostatics evaluated via PME in the program Force Field X. This program was used to 649

re-refine nine mouse and human DNA methyltransferase 1 (Dnmt1) data sets deposited in 650

the Protein databank (PDB). Significant improvements in model quality (presented in 651

Table 1) were achieved as assayed by the MolProbity 241 structure validation tool. The 652

MolProbity score is calibrated to reflect the expected resolution of the X-ray data. After 653

(32)

31

re-refinement, the average MolProbity score was reduced to 2.14, indicating a level of 654

model improvement consistent with collecting data to 0.67 Å higher resolution. For 655

example, the pose of S-adenosyl-L-homocysteine (SAH) from mouse (3PT6) and human 656

(3PTA) structures differed by an RMSD of 1.6 Å before re-refinement, but only 0.9 Å 657

afterwards. 658

Table 1. DNA Methyltransferase 1 (Dnmt1) Models Before and After Polarizable X-Ray 659

Refinement with the Program Force Field X. 660

Protein Databank Re-Refined with Force Field X Statistics MolProbity Statistics MolProbity PDB Res. (Å) R Rfree Score (%) R Rfree Score (%) 3AV4 2.8 0.232 0.267 2.87 68.0 0.238 0.282 2.25 95.0 3AV5 3.3 0.188 0.264 3.09 79.0 0.216 0.275 2.44 97.0 3AV6 3.1 0.195 0.255 2.99 81.0 0.213 0.265 2.37 97.0 3EPZ 2.3 0.213 0.264 2.27 78.0 0.254 0.292 2.09 87.0 3OS5 1.7 0.211 0.238 2.01 54.0 0.182 0.213 1.77 74.0 3PT6 3.0 0.211 0.266 2.95 78.0 0.207 0.268 1.97 99.0 3PT9 2.5 0.196 0.256 2.72 60.0 0.181 0.248 1.90 97.0 3PTA 3.6 0.257 0.291 3.65 57.0 0.211 0.271 2.41 99.0 3SWR 2.5 0.220 0.272 2.69 62.0 0.204 0.264 2.03 95.0 Mean 2.7 0.214 0.264 2.80 68.6 0.212 0.264 2.14 93.3 Mean Improvement 0.67 24.8 661

Figure 1. Polarizable biomolecular X-ray refinement on two Dnmt1 data sets. The left 662

panel shows the deposited pose of SAH from data sets 3PT6 (mouse, grey) and 3PTA 663

(human, cyan) do not agree (coord. RMSD 1.6 Å). In the right panel, the poses of SAH 664

from mouse and human structures are more consistent (coord. RMSD 0.9 Å) after Force 665

(33)

32

Field X refinement.

666

3.8. Prediction of Organic Crystal Structure, Thermodynamics and

667

Solubility

668

It was emphasized in 1998 that predicting crystal structures from chemical composition 669

remained a major unsolved challenge.242 Significant progress has been made since then to 670

address this challenge, as evidenced by successes of the 4th and 5th blind tests of crystal 671

structure prediction (CSP) organized by the Cambridge Crystallographic Data Center 672

(CCDC).243, 244 Prediction of crystal structures is important in the pharmaceutical industry, 673

where extensive experimental screens are necessary to explore the range of stable 674

polymorphs a molecule may form. The unique three-dimensional molecular packing of 675

each polymorph determines its physical properties such as stability and bioavailability. 676

For this reason, both FDA approval and patent protection are awarded to a specific 677

crystal polymorph, rather than to the molecule itself. To illustrate this point, eight 678

companies have filed eleven patents on five possible crystal forms of the molecule 679

cefdinir.245 680

Prediction of thermodynamically stable crystal structures from chemical composition 681

requires a potential energy function capable of distinguishing between large numbers of 682

structures that are closely spaced in thermodynamic stability.246, 247 In this section, we 683

restrict our focus to energy models that explicitly account for electronic polarization 684

classically65, 248, 249 and neglect the more expensive electronic structure methods 685

sometimes used to (re)score favorable structures.250 686

The vast majority of CSP has been limited to using intermolecular potentials that lack 687

explicit inclusion of polarization,249, 251 although its importance has become a topic of 688

(34)

33

interest35, 252-254. Non-polarizable force fields, based on fixed partial charges or fixed 689

atomic multipoles, must implicitly account for the 20% to 40% of the lattice energy 690

attributable induction.249 On the other hand, polarizable models such as the AMOEBA 691

force field for organic molecules 54, 255 based on the Thole damping scheme45 and the 692

Williams-Stone-Misquitta (WSM) method256, 257 for obtaining distributed polarizabilities 693

allow one to include polarization during CSP explicitly. 694

Beyond polarization, modeling the conformational flexibility and corresponding 695

intermolecular energetics of organic molecules via sampling methods such as molecular 696

dynamics is essential to predicting the thermodynamic properties of crystals.258 For 697

example, the structure, stability and solubility of n-alkylamide crystals, from acetamide 698

through octanamide, can be predicted by an alchemical sampling method to compute the 699

sublimation/deposition phase transition free energy.65 700

4. Summary

701

Significant progress has been made in the past decade in developing general-purpose 702

polarizable force fields. Polarizable force fields have exhibited success in disparate 703

research areas including ion solvation, protein-ligand interactions, ion channels and lipids, 704

macromolecular structural refinement and so on. There remain plenty of challenges ahead. 705

The importance of polarization still needs to be established systematically for a wide 706

range of biological systems. While polarizable force fields in principle have better 707

transferability than do non-polarizable force fields, they are also expected to also perform 708

better in a broader range of systems, making parameterization a more elaborate process. 709

In addition to polarization, treatment of other physical effects, including high-order 710

(35)

34

permanent charge distributions interactions, short-range electrostatic penetration and 711

charge-transfer effects need further improvement to advance the overall quality of 712

classical electrostatic models. Because computational efficiency (including the need for 713

parallelization) has been a major barrier to the adoption of polarizable force fields, better 714

and more efficient algorithms are also required to advance the application of polarizable 715

force fields. A future area for advancement is to combine the polarizable force fields with 716

fixed-charge force fields in a multiscale fashion, as is done with QM/MM. Technically 717

this can be achieved straightforwardly but caution is needed to ensure the interactions 718

across the two resolutions are balanced. 719

Acknowledgement.

The authors are grateful to the support provided by Robert A. Welch 720

Foundation (F-1691). 721

(36)

35 REFERENCES

723

1. A. Warshel and M. Levitt, Theoretical Studies of Enzymic Reactions - Dielectric, 724

Electrostatic and Steric Stabilization of Carbonium-Ion in Reaction of Lysozyme. Journal

725

of Molecular Biology, 1976. 103(2): p. 227-249. 726

2. J.A. Mccammon, B.R. Gelin, and M. Karplus, Dynamics of Folded Proteins. Nature, 1977. 727

267(5612): p. 585-590. 728

3. M.A. Spackman, The Use of the Promolecular Charge Density to Approximate the 729

Penetration Contribution to Intermolecular Electrostatic Energies. Chemical Physics

730

Letters, 2006. 418(1-3): p. 158-162. 731

4. D. Nachtigallova, P. Hobza, and V. Spirko, Assigning the Nh Stretches of the Guanine 732

Tautomers Using Adiabatic Separation: CCSD(T) Benchmark Calculations. Journal of

733

Physical Chemistry A, 2008. 112(9): p. 1854-1856. 734

5. W.D. Cornell, P. Cieplak, C.I. Bayly, I.R. Gould, K.M. Merz, D.M. Ferguson, D.C. 735

Spellmeyer, T. Fox, J.W. Caldwell, and P.A. Kollman, A 2nd Generation Force-Field for the 736

Simulation of Proteins, Nucleic-Acids, and Organic-Molecules. Journal of the American

737

Chemical Society, 1995. 117(19): p. 5179-5197. 738

6. A.D. MacKerell, D. Bashford, M. Bellott, R.L. Dunbrack, J.D. Evanseck, M.J. Field, S. 739

Fischer, J. Gao, H. Guo, S. Ha, D. Joseph-McCarthy, L. Kuchnir, K. Kuczera, F.T.K. Lau, C. 740

Mattos, S. Michnick, T. Ngo, D.T. Nguyen, B. Prodhom, W.E. Reiher, B. Roux, M. 741

Schlenkrich, J.C. Smith, R. Stote, J. Straub, M. Watanabe, J. Wiorkiewicz-Kuczera, D. Yin, 742

and M. Karplus, All-Atom Empirical Potential for Molecular Modeling and Dynamics 743

Studies of Proteins. Journal of Physical Chemistry B, 1998. 102(18): p. 3586-3616.

744

7. H. Valdes, K. Pluhackova, M. Pitonak, J. Rezac, and P. Hobza, Benchmark Database on 745

Isolated Small Peptides Containing an Aromatic Side Chain: Comparison between Wave

(37)

36

Function and Density Functional Theory Methods and Empirical Force Field. Physical

747

Chemistry Chemical Physics, 2008. 10(19): p. 2747-2757. 748

8. W.L. Jorgensen, D.S. Maxwell, and J. Tirado-Rives, Development and Testing of the OPLS 749

All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids.

750

Journal of the American Chemical Society, 1996. 118(45): p. 11225-11236. 751

9. J.W. Ponder and D.A. Case, Force Fields for Protein Simulations. Advances in Protein 752

Chemistry, 2003. 66: p. 27-85. 753

10. J. Rezac, P. Jurecka, K.E. Riley, J. Cerny, H. Valdes, K. Pluhackova, K. Berka, T. Rezac, M. 754

Pitonak, J. Vondrasek, and P. Hobza, Quantum Chemical Benchmark Energy and 755

Geometry Database for Molecular Clusters and Complex Molecular Systems

756

(Www.Begdb.Com): A Users Manual and Examples. Collection of Czechoslovak Chemical

757

Communications, 2008. 73(10): p. 1261-1270. 758

11. J. Rezac, K.E. Riley, and P. Hobza, S66: A Well-Balanced Database of Benchmark 759

Interaction Energies Relevant to Biomolecular Structures. Journal of Chemical Theory

760

and Computation, 2011. 7(8): p. 2427-2438. 761

12. K. Berka, R. Laskowski, K.E. Riley, P. Hobza, and J.i. Vondrášek, Representative Amino 762

Acid Side Chain Interactions in Proteins. A Comparison of Highly Accurate Correlated Ab

763

Initio Quantum Chemical and Empirical Potential Procedures. Journal of Chemical Theory

764

and Computation, 2009. 5(4): p. 982-992. 765

13. J.A. Barker, Statistical Mechanics of Interacting Dipoles. Proceedings of the Royal Society 766

of London. Series A, Mathematical and Physical Sciences, 1953. 219(1138): p. 367-372. 767

14. J.W. Caldwell and P.A. Kollman, Structure and Properties of Neat Liquids Using 768

Nonadditive Molecular Dynamics: Water, Methanol, and N-Methylacetamide. Journal of

769

Physical Chemistry, 1995. 99: p. 6208-6219. 770

Figure

Table 1.  DNA Methyltransferase 1 (Dnmt1) Models Before and After Polarizable X-Ray 659

Références

Documents relatifs

Molecular dynamics simulations of ions in a water drop were carried out with the AMOEBA polarizable force field to (i) charac- terize their first two solvation shells, and (ii)

Here we enable the direct construction of flexible molecular force fields from high-level ab initio calculations by incorporating spatial and temporal physical symmetries into

In point-charge force fields, the partial charges are usually determined by fitting to quantum mechanical (QM) electrostatic potential (ESP) or interaction energy with water

Using conservation of energy — a fundamental property of closed classical and quantum mechanical systems — we develop an efficient gradient-domain machine learning (GDML) approach

In the present work, we assess the validity of elastic models for the shear viscosity and the α-relaxation time of supercooled water, using molecular dynamics simulations with

Our vision-based method combin- ing numerical simulation and Neural Networks, predicts the range of the force applied by a needle using only 2D images of the scleral deformation

However, the volumes of the boron reference units are roughly the same (Table IX) which means that the structural swelling involving boron units are essentially due to the NBO

Characterising the lowest possible error within a given body order and polynomial degree for individual fits is helpful when thinking about the errors of a general force field